Coordination of frontline defense mechanisms under ... - BioMedSearch

0 downloads 0 Views 3MB Size Report
Jul 27, 2010 - fepC. fhuG bdb. sojC2. gabT. 2562H. 0053H. 0058H. 0248C. 2299H. 2563H ..... lumped into a loosely defined 'general stress response.'.
Molecular Systems Biology 6; Article number 393; doi:10.1038/msb.2010.50 Citation: Molecular Systems Biology 6:393 & 2010 EMBO and Macmillan Publishers Limited All rights reserved 1744-4292/10 www.molecularsystemsbiology.com

Coordination of frontline defense mechanisms under severe oxidative stress Amardeep Kaur1, Phu T Van1, Courtney R Busch2, Courtney K Robinson2, Min Pan1, Wyming Lee Pang1, David J Reiss1, Jocelyne DiRuggiero2 and Nitin S Baliga1,* 1

Institute for Systems Biology, Seattle, WA, USA and 2 Department of Biology, Johns Hopkins University, Baltimore, MD, USA * Corresponding author. Baliga Lab, Institute for Systems Biology, 1441 North 34th Street, Seattle, WA 98103, USA. Tel.: þ 1 20 67 32 1266; Fax: þ 1 20 67 32 1299; E-mail: [email protected] Received 6.10.09; accepted 31.5.10

Complexity of cellular response to oxidative stress (OS) stems from its wide-ranging damage to nucleic acids, proteins, carbohydrates, and lipids. We have constructed a systems model of OS response (OSR) for Halobacterium salinarum NRC-1 in an attempt to understand the architecture of its regulatory network that coordinates this complex response. This has revealed a multi-tiered OS-management program to transcriptionally coordinate three peroxidase/catalase enzymes, two superoxide dismutases, production of rhodopsins, carotenoids and gas vesicles, metal trafficking, and various other aspects of metabolism. Through experimental validation of interactions within the OSR regulatory network, we show that despite their inability to directly sense reactive oxygen species, general transcription factors have an important function in coordinating this response. Remarkably, a significant fraction of this OSR was accurately recapitulated by a model that was earlier constructed from cellular responses to diverse environmental perturbations—this constitutes the general stress response component. Notwithstanding this observation, comparison of the two models has identified the coordination of frontline defense and repair systems by regulatory mechanisms that are triggered uniquely by severe OS and not by other environmental stressors, including sub-inhibitory levels of redox-active metals, extreme changes in oxygen tension, and a sub-lethal dose of c rays. Molecular Systems Biology 6: 393; published online 27 July 2010; doi:10.1038/msb.2010.50 Subject Categories: metabolic and regulatory networks; microbiology and pathogens Keywords: gene regulatory network; microbiology; oxidative stress This is an open-access article distributed under the terms of the Creative Commons Attribution Noncommercial Share Alike 3.0 Unported License, which allows readers to alter, transform, or build upon the article and then distribute the resulting work under the same or similar license to this one. The work must be attributed back to the original author and commercial use is not permitted without specific permission.

Introduction Reactive oxygen species (ROS), such as hydrogen peroxide  (H2O2), superoxide (O 2 ), and hydroxyl (OH ) radicals are normal by-products of aerobic metabolism. Evolutionarily conserved mechanisms including detoxification enzymes (peroxidase/catalase and superoxide dismutase (SOD)) and free radical scavengers manage this endogenous production of ROS. Oxidative stress (OS) is a condition reached when certain environmental stresses or genetic defects cause the production of ROS to exceed the management capacity. The damage to diverse cellular components including DNA, proteins, lipids, and carbohydrates resulting from OS (Imlay, 2003; Apel and Hirt, 2004; Perrone et al, 2008) is recognized as an important player in many diseases and in the aging process (Finkel, 2005). A great deal of information exists on specific detoxification enzymes, ion scavengers, and their associated regulators. For & 2010 EMBO and Macmillan Publishers Limited

instance, in Eshcherichia coli, it is known that the oxyR regulon is induced by higher levels of H2O2, whereas the soxRS regulon is activated by exposure to redox-cycling agents (Zheng and Storz, 2000; Green and Paget, 2004; Imlay, 2008). Activation of many of these regulators by metals, radiation, and starvation conditions shows how diverse environmental factors (EFs) can directly or indirectly increase production of ROS. Not surprisingly, the regulation of OS response (OSR) is tied to circuits that manage diverse processes including metal trafficking, nutrient transport, and general aspects of metabolism. The function of metal trafficking is especially important given the function of transition metals, iron in particular, in the production of highly reactive OH through the Fenton reaction. It is clear from all of these earlier studies that a systems approach will be necessary to fully appreciate the highly interconnected nature of the regulatory networks that manage OSR (Temple et al, 2005). Molecular Systems Biology 2010 1

Systems model of oxidative stress response A Kaur et al

Given that many archaea thrive in extreme environments associated with elevated production of ROS, it is intriguing to consider these organisms as models for characterizing global regulatory networks that efficiently manage extreme OS during normal metabolism (Joshi and Dennis, 1993; Kawakami et al, 2004; Limauro et al, 2008). Furthermore, although archaeal transcriptional regulation is mediated by regulators of bacterial ancestry, many of the known bacterial OSR regulators are absent in this domain of life (Coulson et al, 2007). Here, we have applied a systems approach to characterize the OSR of an archaeal model organism, Halobacterium salinarum NRC-1. This haloarchaeon grows aerobically at 4.3 M salt concentration in which it routinely faces cycles of desiccation and rehydration, and increased ultraviolet radiation—both of which can increase the production of ROS (Farr and Kogoma, 1991; Oliver et al, 2001). We have reconstructed the physiological adjustments associated with management of excessive OS through the analysis of global dynamic transcriptional changes elicited by constant exposure to growth subinhibitory and sub-lethal levels of H2O2 and paraquat (PQ—a redox-cycling drug that produces O 2 ; Hassan and Fridovich, 1979) as well as during subsequent recovery from these stresses. We have integrated all of these data into a unified model for OSR to these ROS to discover conditional functional links between protective mechanisms and normal aspects of metabolism. Subsequent phenotypic analysis of gene deletion strains has verified the conditional detoxification functions of three putative peroxidase/catalase enzymes, two SODs, and the protective function of rhodopsins under increased levels of H2O2 and PQ. Similarly, we have also validated ROS scavenging by carotenoids and flotation by gas vesicles as secondary mechanisms that may minimize OS. Although it has been known that OS is a component of diverse environmental stress conditions, we quantitatively show for the first time that much of the transcriptional responses induced by the two treatments could indeed have been predicted using a model constructed from the analysis of transcriptional responses to changes in other EFs (UV and g-radiation, light, oxygen, and six metals). However, using specific examples, we also reveal the specific components of the OSR that are triggered only under severe OS. Notably, this model of OSR gives a unified perspective of the interconnections among all of these generalized and OSspecific regulatory mechanisms.

Results and discussion We present the results and discussion in six sections (A–F) starting with (A) characterization of growth and survival characteristics of H. salinarum NRC-1 under increasing doses of H2O2 and PQ, (B) analysis of global dynamic transcriptional responses to selected sub-inhibitory and sub-lethal doses of the two stressors, (C) inference of a globally integrated model of conditionally co-regulated genes and their sub-circuits, (D) experimental validation of important hypotheses regarding primary and secondary defense mechanisms and their functional relationships, (E) statistical evaluation of generalized and specific aspects of OS management, and (F) summary of the integrated program for OS management. 2 Molecular Systems Biology 2010

Distinct differences in phenotypic responses to H2O2 and PQ We first characterized survival and growth characteristics of H. salinarum NRC-1 with increasing doses of H2O2 and O 2. Although peroxide stress was generated by directly adding H2O2 to a desired concentration (titration doses for survival studies: 0–40 mM; growth studies: 0–7 mM), O 2 radicals were generated indirectly during metabolism of exogenously added PQ (titration doses for survival studies: 0–10 mM; growth studies: 0–0.5 mM) (Hassan and Fridovich, 1979) (Figure 1). Interestingly, the cells withstood exogenously added H2O2 stress up to a concentration of B30 mM beyond which their survival dropped precipitously. This sharp transition was also observed at the level of growth wherein the cells grew normally in up to 5–6 mM H2O2 and a smallest increase thereafter resulted in complete growth inhibition. This indicates that in the absence of earlier conditioning to low doses of OS (Cabiscol et al, 2000; Brioukhanov et al, 2006; Limauro et al, 2006), there exists a threshold to the cellular capacity to detoxify H2O2 and that even traces of peroxide beyond this threshold can result in devastating oxidative damage to cellular components. This phenotypic behavior under H2O2 stress is in stark contrast to the gradual loss in survival and growth observed under increasing O 2 stress. The growth and survival assays aided in the selection of appropriate concentrations of H2O2 and PQ for subsequent transcriptomic analysis. The enormous amount of the two stress agents required to achieve sub-lethal effects is a testament to the remarkable capacity of this organism to efficiently eliminate and keep intracellular concentration of ROS within physiologically acceptable limits.

Global transcriptional analysis reveals shared and specific responses to peroxide and O 2 We sought to model OSR at a systems scale by characterizing OS-management mechanisms and unraveling the gene regulatory programs that coordinate their operation with other aspects of metabolism. We approached this problem by charting dynamic global transcriptional changes for up to 4 h during and after exposure to sub-lethal concentrations of each PQ (4 mM) and H2O2 (25 mM) that resulted in 20–30% loss of survival (Figure 2; Supplementary Figures S1 and S2; Supplementary Tables S1–S3). Further, we also characterized transcriptional responses during nearly 6 h of acclimation to a growth sub-inhibitory dose of PQ (0.25 mM) (this was not possible for peroxide stress owing to the sharp transition in growth with small concentration increases of H2O2) (Supplementary Figure S3; Supplementary Table S4). Together with control experiments and replicates, this represented nearly 300 high-density microarray experiments. We summarize discoveries from integrated analyses of data from these experiments (Materials and methods) (Supplementary Figure S1). Exposure to both ROS resulted in differential regulation of a larger proportion of genes over time involving up to 50% of all genes at steady state. In contrast, transcriptional changes became somewhat subdued over time during recovery from both OS conditions (Supplementary Figure S1). Nearly half of & 2010 EMBO and Macmillan Publishers Limited

Systems model of oxidative stress response A Kaur et al

A

H2O2

C

0 mM 4 mM 5 mM 6 mM 7 mM

OD600

Survival (Nt /N0)

1

H2O2 10

1

0.1

0.01

0.1 0

10

20

30

40

50

0

10

20

Time (h)

B

D

PQ

1

30

40

50

8

10

(mM)

OD600

Survival (Nt /N0)

0 mM 0.125 mM 0.25 mM 0.5 mM

PQ

10

1

0.1

0.01

0.1 0

10

20

30

40

50

Time (h)

0

2

4

6 (mM)

Figure 1 Growth and survival of H. salinarum NRC-1 with H2O2 and PQ. Effect of increasing concentrations of H2O2 and PQ on growth (A, B) and survival (C, D) of H. salinarum NRC-1 in complete media (CM). Growth was measured by determining increase in cell density (OD600) at 30 min intervals using a Bioscreen instrument. Survival was calculated as ratio of number of viable cells (N) after 2 h treatment with varying doses of H2O2 or PQ to number of viable cells (No) in control. Black arrows represent dosage at which microarray analyses were performed. Data shown is representative of the experiments, which were repeated three times.

all differentially regulated genes (563 genes) were shared by the two stress responses—perhaps reflecting a generalized OS component (Supplementary Table S1). Notably, differential regulation of an equivalent fraction of genes (H2O2: 366 genes; PQ: 536 genes) was unique to each response. Some unique aspects of responses to the two ROS and between responses to sub-lethal and sub-inhibitory doses of PQ are noted in the Supplementary information (Figure 2; Supplementary Figures S2 and S3). By simultaneously analyzing putative functional assignments (Bonneau et al, 2004) and transcriptional changes, we have reconstructed physiological consequences of long-term exposures to and recovery from excessive H2O2 and O 2 stress (Figure 2; Supplementary Figures S1–S3; Supplementary Tables S2–S4). We provide a summary of this analysis with emphasis on OS-related functions; observations pertaining to adjustments in general aspects of physiology including ribosome biogenesis, oxidative phosphorylation, and carbohydrate and nucleotide metabolism during the OSR are summarized in the Supplementary information.

upregulated by both treatments (Figure 2O, P and S; Supplementary Table S2). Notably, in addition to upregulation of the classic peroxidase/catalase [VNG6294G (perA)— PF00141], we observed transcriptional induction of at least two additional enzymes of related function—VNG0798H, a putative dye-decolorizing peroxidase (PF04261, COG2837) (Sugano et al, 2007; Scheibner et al, 2008) and VNG0018H, whose predicted structure matches that of catalase [PDB: 1cf9A2; (Mate et al, 1999)]. VNG1190G (sod1) and peroxiredoxins [VNG1197G (bcp), VNG2311H] were upregulated to a higher degree with PQ than with H2O2, suggesting that different regulatory circuits are responsible for managing ROS generated by the two treatments (Brioukhanov et al, 2006; Limauro et al, 2006, 2008). Upregulation of a number of other processes was also consistent with their functions in relation to OS including carotenoid synthesis [VNG1458G (crtB1), VNG1755G (crtI2)] and glycerol metabolism [VNG6277G (ugpB), VNG6279G (ugpA), VNG1967G (glpK), VNG6270G (gldA)] (Pahlman et al, 2001) (Figure 2A, B, and N).

ROS scavenging and detoxification

Repair of oxidative damage

Many of the primary OS-management systems including detoxification by SODs, peroxidase/catalase, and peroxiredoxins, and radical scavenging by antioxidants, and reduction of radicals by redoxins (Storz and Imlay, 1999; Grant, 2001; Zheng et al, 2001; Toledano et al, 2007; Imlay, 2008) were

Although DNA repair mechanisms were in general regulated in a similar manner by both agents (Figure 2T; Supplementary Table S2), PQ treatment uniquely upregulated certain repair and recombination genes including VNG2473G (radA1), ssDNA binding (VNG1255C and VNG 2160C), and DNA

& 2010 EMBO and Macmillan Publishers Limited

Molecular Systems Biology 2010 3

Systems model of oxidative stress response A Kaur et al

I H2O2

H2O2

PQ

A

II H2O2

PQ

H

Q

I

R

H2O2

PQ

V

H2O2

PQ

PQ

W

B

J C

D

– 0.3

E

I

F K

S

L

M G

T

N

0.3

A B C D E F&G H I J K L M N O P Q R S T U

ABC transporters Lipid metab Glycolysis TCA cycle RNA plymerase Nucleotide metab OxPhos Cobalamin biosynth Ribosomal proteins DMSO metabolism Mo cofactor biosynth Phototrophy Carotenoid biosynth Trx/Grx Detoxification Gas vesicle biogenesis AA metabolism Fe metabolism Repair genes Proteases

II Unique groups

O P

0

Function class

V W

U

SAM — Group 1 SAM — Group 2

Experimental design 80 5 20 –1 10 40 160 Time (min) 25 mM H2O2 or 4 mM PQ

Figure 2 Temporal transcriptome changes in H. salinarum NRC-1 treated with sub-lethal doses of H2O2 (25 mM) or PQ (4 mM). Relative changes in transcription of all genes were measured immediately before and up to 160 min after adding 25 mM H2O2 or 4 mM PQ. The mRNA log10 ratios were normalized to the first condition (control) and data were filtered (log10 ratio 40.05, l¼15 in at least one condition; see Materials and methods for details). I. Differentially regulated genes are grouped into different functional groups (A–U) (see Supplementary Table S2 for details). II. Transcriptional profiles that distinguished OSRs to H2O2 and PQ (Groups V and W) were identified using the significance analysis of microarrays (SAM) algorithm (Tusher et al, 2001).

4 Molecular Systems Biology 2010

& 2010 EMBO and Macmillan Publishers Limited

Systems model of oxidative stress response A Kaur et al

helicase [VNG1406G (rhl)]. Although it is clear from upregulation of proteases [VNG0205H, VNG0354C, VNG2096G (cctB), VNG0226G (htrA), VNG1951G (sub), VNG2323H, VNG6201G (hsp5), VNG6361G (npa)] that protein degradation was upregulated by both treatments (Figure 2U), it was striking that unlike the upregulation of ribosomal genes for replacing damaged proteins (Thorpe et al, 2004) under H2O2 stress, these genes were downregulated with PQ—even after the stress was taken away. We speculate that the sub-lethal dose of PQ (4 mM) was more damaging to proteins relative to H2O2 (25 mM), despite their comparable phenotypic effects (20–30% loss in survival). This is not surprising as PQ primarily generates O 2 ions, which are known to be particularly deleterious to proteins as they increase the production of highly reactive OH radicals through the Fenton reaction by releasing free Fe2 þ from solvent-exposed 4Fe-4S cluster of labile dehydratases (Imlay, 2003). As expected, lowering the dose of PQ to sub-inhibitory levels induced transcription of both proteases and ribosome biogenesis (Supplementary Figure S3).

An integrated predictive model for OS management Data presented here and earlier (Causton et al, 2001) have clearly shown that OSR is a global phenomenon impacting a wide array of cellular processes encoded by at least 50% of all genes (Figure 2; Supplementary Figure S2 and S3; Supplementary Table S1). We have taken a step toward construction of a predictive model for global transcriptional coordination of the OSR. Briefly, using the cMonkey algorithm (Reiss et al, 2006), we integrated transcriptome changes, gene functional associations (phylogenetic profiles, chromosomal proximity, operon organization, metabolic networks, etc.), and de novo discovered conserved cis-regulatory motifs to identify subsets of genes that are conditionally co-regulated by OS (biclusters). Subsequently, we constructed an environment and gene regulatory influence network for OS (EGRINos) by performing linear regression and model selection using the Inferelator algorithm (Bonneau et al, 2006, 2007). Analysis of regulatory influences of TFs and the two OS agents on each bicluster within this model has recapitulated earlier known regulatory phenomena for managing OS and revealed new putative regulatory links among diverse cellular processes.

Regulation of Fe–S metabolism Most transcriptional changes induced by both treatments were over-represented for Fe-dependent and Fe–S-containing enzymes (Supplementary Figure S4; Supplementary Tables S5 and S6). For instance, a simple analysis of Fe–S cluster-associated motif-containing proteins shows that six of eight (P¼0.1) putative electron transfer system proteins (CX2CX2CX3C) and seven of nine (P¼0.08) putative enzymes requiring S-adenosylmethionine (SAM) (CX3CX2C) were differentially expressed under OS (Supplementary Tables S5 and S6). This is not surprising as Fe–S clusters are used as cofactors by many redox proteins, radical SAM enzymes, and RNA polymerase (Boyd et al, 2009), and are important for electron transfer reactions and gene regulation as they sense changes in Fe, O2, and O 2 (Fontecave, 2006; Outten, 2007) (Figures 2, 3; Supplementary Tables S2 and S3). It has been proposed that conversion of Fe2 þ to Fe3 þ could trigger three types of responses: a Fe-deficiency such as response to bring in more Fe into the cell; an effort to repair or re-synthesize proteins with damaged Fe–S clusters and S-containing amino acids; and a protective response to sequester Fe in a non-reactive form (Varghese et al, 2007; Imlay, 2008). Consistent with this, cellular effort to scavenge and sequester Fe, and simultaneously repair both damaged Fe–S clusters and S-containing amino acids was reflected in the upregulation of genes encoding Fe-scavengers: [putative siderophore biosynthesis [VNG6213G (iucB), VNG6212G (iucA), VNG6214G (hxyA), VNG6216G (iucC)] (Imlay, 2006)]; iron sequestration [ferritin [VNG2443G (dpsA)] (Andrews et al, 2003; Reindel et al, 2006; Shukla, 2006)]; S uptake [VNG1592G (cysT2)]; assembly of Fe–S clusters [VNG0524G (yurY), VNG0525C, VNG0527C] (Takahashi and Tokumoto, 2002; Loiseau et al, 2003); biosynthesis of cysteine and methionine [VNG0796G (cgs), VNG2420G (metA), VNG2421G (hal), and VNG1301G (cysK)]; and repair of oxidized cysteines and methionines [VNG1180G (msrA)] (Stadtman et al, 2003; Stadtman, 2006; Metayer et al, 2008) (Figure 2; Supplementary Table S2). & 2010 EMBO and Macmillan Publishers Limited

Discovery of co-regulated genes reveals new regulatory links among different OS components Altogether, 1165 genes were grouped into 100 biclusters using the cMonkey algorithm (Supplementary Table S7). Using two examples, we show how this integrated the analysis of genes, environmental conditions, functional associations, and shared cis-regulatory motifs within individual biclusters has discovered numerous regulons and provided evidence for conditional co-regulation of diverse functions during OSR (Figure 3). Example 1: Analysis of bicluster #84 (bc84) has provided evidence for the potential transcriptional co-regulation of sod1 with several redoxins, a thioredoxin reductase, bacterioferritin, cysteine biosynthesis/repair enzymes, molecular chaperones, and five genes with no matches to protein sequence or signature of characterized function. Though it has been shown that sod1 and cysteine repair systems are part of the SoxRS regulon and Fe scavenging is part of the OxyR regulon (Imlay, 2008), global analysis has shown the two regulons to be overlapping (Farr and Kogoma, 1991; Zheng et al, 2001). Similarly, the gene composition of bc84 suggests that this regulon is analogous to a functional hybrid of the SoxRS and OxyR systems in E. coli. However, there are no orthologs for either of these regulators in H. salinarum NRC-1 (Coulson et al, 2007). Subsequent regulatory network inference with Inferelator has helped identify potential regulator(s) that mediate this control (see below). Notably, co-expression of the 23 genes was especially prominent in PQ-treated cells. Although this is consistent with the O 2 -specific function of some of these genes (e.g. sod1), it also suggests similar ROS-specific functions for some of the other genes. Example 2: In bc12, we have discovered potential coregulation of genes encoding siderophore biosynthesis, ferrichrome transport, a thioredoxin reductase, and eight earlier unknown function proteins. The regulatory link Molecular Systems Biology 2010 5

Systems model of oxidative stress response A Kaur et al

A

4

Bicluster 84 Superoxide dismutase [Mn] 1 Thioredoxin reductase Glutaredoxin Putative glutaredoxin Putative peroxidoxin Putative glutaredoxin Oxidoreductase Quinone oxidoreductase Dehydrogenase/oxidoreductase Cysteine synthase Cysteine protease Methionine sulfoxide reductase Bacterioferritin Molecular chaperone Universal stress protein Cobalamin adenosyltransferase Transcription regulator Unknown function Molecular chaperone Unknown function Unknown function Unknown function Unknown function

Bicluster average Constant stress Recovery

Paraquat

3 Normalized expression

sod1 trxB2 2115H 2310H 2311H 1546H yqjM yfmJ nolA cysK 0557H msrA dpsA 0283C 2520C 1547C 2163H 0282H 1047H 0360C 1092C 2162C 2432C

H2O2

2 1 0 –1 –2 0

20

40 Conditions

1 3 5 7 9 11 13 15 17 19 21 Motif 1: E = 2.8e–08

1

60

3

5

80

7 9 11 13 15 Motif 2: E = 0.035

B 4

Bicluster 12 Siderophore biosynthesis operon Monooxygenase Thioredoxin reductase Ferrichrome ABC transporter Ferrichrome ABC transporter Ferric enterobactin transporter Ferrichrome ABC transporter Diaminobutyrate decarboxylase Spo0A activation inhibitor Aminobutyrate aminotransferase Periplasmic binding protein Unknown function Unknown function Unknown function Unknown function Unknown function Unknown function Unknown function Unknown function

H2O2

Paraquat Bicluster average Constant stress Stress recovery

3 Normalized expression

iucABC hxyA txrB3 yfmD2 yfmF fepC fhuG bdb sojC2 gabT 2562H 0053H 0058H 0248C 2299H 2563H 6152H 6159H 6160H

2 1 0 –1 –2 0

20

40

60

80

Conditions fhuG fepC

yfmF

yfmD2

bdb iucA iucB hxyA iucC

1 3 5 7 9 11 13 15 17 19 21

Operon

Motif: E = 3.4e–19

Figure 3 Discovery of conditionally co-regulated genes reveals novel physiological links. Conditionally co-regulated genes were discovered using the cMonkey algorithm (see text for details). (A, B) Shows two example biclusters: bc84 (23 genes) and bc12 (21 genes), respectively. The plots show mean and variance-normalized transcriptional changes (log10 ratios) for all genes (gray) in each bicluster with an overlaid averaged profile (black). Filled triangles on x axis indicate time course analysis of responses during continued exposure to 25 mM H2O2 or 4 mM PQ. The open triangles, on the other hand, indicate time course analysis of recovery from 2 h exposure to 25 mM H2O2 or 4 mM PQ. Genes and their putative functions are indicated in the table. Sequence logos for the conserved motifs are shown in lower right of each panel, with numbers indicating nucleotide position and size of letters indicating relative conservation of bases. Genes possessing conserved DNA sequence motifs are marked with squares, triangles, or circles. (B) Lower right: operon-like organizations of nine genes within bicluster 12.

between siderophore biosynthesis and ferrichrome transport is not surprising given their shared functions in managing intracellular Fe pool (Andrews et al, 2003), and their conditional co-regulation during an OSR, especially H2O2 stress, in E. coli and Bacillus subtilis (Imlay, 2008). However, evidence for co-regulation of this process with a chromosome partitioning gene [VNG6153G (sojC2)], thioredoxin reductase [VNG2301G (trxB3)], and eight genes of unassigned functions is new information. A single conserved palindromic motif within the promoters of all genes within this bicluster 6 Molecular Systems Biology 2010

suggested that this regulon is under the direct control of a single TF (Figure 3). In comparing the distribution of genes in the two biclusters discussed above, we also note that this integrated analysis has helped hypothesize condition-specific functions for genes of related function [redox homeostasis (e.g. thioredoxin reductase, glutaredoxin, and peroxiredoxins), 12 Fe-management genes, proteolysis, cobalamin biosynthesis, and O 2 metabolism] as well as putative functions for at least three genes (VNG1092C, VNG0248C, VNG2299H) of earlier unknown & 2010 EMBO and Macmillan Publishers Limited

Systems model of oxidative stress response A Kaur et al

function. Interactive exploration of all biclusters to enable similar discoveries is possible online at http://baliga. systemsbiology.net/drupal/content/egrin-oxidative-stress.

A unified model for global transcriptional coordination of responses to treatments with H2O2 and PQ OSR regulators are typically associated with redox-active cofactors such as haem, flavins, pyridine nucleotides, Fe–S clusters, and redox-sensitive amino-acid side chains such as cysteinyl thiols (Green and Paget, 2004). Altogether, 26 TFs and 17 signal transducers [containing methyl-accepting chemotaxis protein (MCP)-signaling domain] in H. salinarum NRC-1 have matches to motif signatures for these OS-sensing features (Zheng and Storz, 2000; Fomenko and Gladyshev, 2002; Paget and Buttner, 2003; Fontecave, 2006). A significant fraction of these regulators (21 TFs and 13 signal transducers) was differentially regulated in response to H2O2 and/or PQ treatments (Supplementary Table S5). Although the high degree of cross-regulation that we observed might result from chemical inter-conversions among ROS (Liochev and Fridovich, 2007), it is more likely because of regulators that sense a globally affecting parameter such as the overall redox status (Demple, 1997; Zheng and Storz, 2000) or metal ion composition of the cell (Touati, 2000; Moore and Helmann, 2005; Lee and Helmann, 2006). However, even though the trigger (e.g. redox status) for many of these regulators is the same, their mechanism of activation and the genes they regulate are nonetheless different. In E. coli, for instance, in response to O 2 stress, SoxR gets activated through oxidation of its Fe–S cluster and regulates a specific subset of genes that is distinct from those controlled by OxyR, which is activated through oxidation of cysteine residue(s) under H2O2 stress (Zheng and Storz, 2000; Imlay, 2008). In our study as well, we have observed unique transcriptional responses, such as upregulation of three thioredoxins [VNG1012H, VNG1259G (trxB2), VNG2600G (trxA2)] and a DNA repair gene VNG2473G (radA1) by PQ, but not by H2O2, suggesting involvement of several ROS-specific regulatory mechanisms in mediating the global OSR. We have taken a step toward a systems level model (Figure 4) that collectively explains these OS-induced gene-expression changes as a function of shared and unique influences of TFs and two ROS. These regulatory influences in EGRINOS recapitulate the global transcriptional changes observed during an OSR (Supplementary Figure S5). EGRINOS also provides meaningful insight into the interrelationships among stress-management systems and general aspects of physiology under H2O2 and PQ stress (Figure 4). The architecture of the model reveals that the OSR network has significant overlap with response networks that manage other kinds of stress. For instance, a common set of TFs are predicted to coordinate genes in bc6, which contains the putative dye-decolorizing peroxidase (VNG0798H), with those in bc84 (contains sod1) and bc12 (contains Fe-management proteins) (bc84 and bc12 are discussed in detail above). Likewise, OSR genes within bc26 (a monooxygenase, several thioredoxin reductases, methionine sulfoxide reductase msrA, cysteine synthetase cysK, and a putative protease) are positively influenced by the same putative TF (Imd1) that also negatively influences & 2010 EMBO and Macmillan Publishers Limited

transcription of cobalamin synthesis and ATP synthase genes in bc91. Finally, three biclusters containing genes of diverse redox-reactive functions (bc17—carotenoid biosynthesis; bc33—siderophore biosynthesis; bc71—antioxidant synthesis) are co-regulated by TFs with redox-sensing and metalbinding domains. In the following sections, we show how EGRINOS can be used as a framework to formulate experimentally testable hypotheses for characterizing general and specific components of the OSR. The architecture of this statistically inferred network (Figure 4A) can be used to formulate hypotheses regarding combinatorial control of cellular processes during an environmental response, OSR in this case. These hypotheses can be tested by validating physical interactions of TFs with promoters of genes they are predicted to regulate, and by analyzing transcriptional and phenotypic consequences of genetically perturbing TFs in the network. Using a combination of such experimental validations, we have overcome the complexities presented by extensive buffering within such networks because of multiple redundancies for regulation and stress management. We have validated regulatory influences of six TFs [Trh4 and five general transcription factors (GTFs— TFBb, c, d, f, and g)] in the OSR network by showing their physical interactions with promoters of target genes, and consequence of deleting some of these TFs on expression of these genes (TFBd and c) and, more importantly, on survival rate (TFBc) under OS. Specifically, we have validated combinatorial control of genes within the same bicluster (bc88) by multiple TFs (Trh4, TFBb, TFBf, and TFBg; Figure 4B; Supplementary Table S9), differential regulation of genes in different biclusters (bc3, bc81, bc72, and bc33) by the same TF (TFBd) (Figure 4C), and diminished capacity to withstand OS on deleting a predicted TF (TFBc) within this network (Figure 4D). The consequence of deleting TFBc on increased susceptibility to OS can now be further investigated in the context of its function in coordinating specific processes as predicted by EGRINOS.

Phenotypic analysis of mutants reveals a multi-tiered program for OS management In addition to validating architecture of EGRINOS, we have also conducted phenotypic analysis of strains with in-frame chromosomal deletions in genes identified to be important in OS management based on their putative functions, their transcriptional changes in response to OS, and the EGRINos model. Specifically, we investigated the functions of two SODs, three putative peroxidase/catalase genes, carotenoids, rhodopsins, and gas vesicles in OS management (Figures 5 and 6).

Two SODs (SOD1 and SOD2) offer protection against PQ treatment but not against H2O2 According to the EGRINos model, the two SOD genes are not co-regulated even by PQ treatment, which predominantly generates O 2 stress in which their function is expected to be most relevant, suggesting distinct functions for the two orthologs. The co-regulation of VNG1190G (sod1) with genes encoding methionine sulfoxide reductase [VNG1180G Molecular Systems Biology 2010 7

Systems model of oxidative stress response A Kaur et al

A

bc72 Ribosome oxidoreductases

TF15 TF29

TF6

5050H

9 10

TF17

88 ABC transporters dehydrogenases

TF19

sirR

14

trh2 trh6 cinR 1405C 1510C

15 17 19

tbpC tfeA boa3 0142C gvpE1 tbpE tfbG tfbB 0066H 0751C 2522C 5075C

0293H

TF21 nusG

sirR 9 PO4transporters

7

TF8

1488G TF9

5 6

TF5

TF7

46

TF2 2

sod2 dehydrogenases

TF26

12

21 22 23

nirH 2476C tfbF trh7 5144H cspD1 nusG imd2 snp 0258H

24

tfbA tfbD tfbE tbpD dmsR gvpE2 0462C 1179C 1237C 2112C 6143H 6193H 6387H asnC troR 0471C 2268H trh3 pai1

Fe metabolism 84

sod1 msrA dpsA trxB2

22

Purine metabolism

TF4

Legend

26 28

0.1

29

2163H msrA Glutaredoxin oxidoreductases

TF28 91 cobalamin syn. ox. phosph.

0389C

26

imd1

B

0.2 0.4 Activates Represses Combines

D

C TF17

TF22

bc12

TF8

TF24 P-value 0.0011 0.0559

bc33

bc3 bc81

0.3

bc88

bc59

bc25 TF19

log10 ratios

0.2 0.1

bc72

0511H 0988H 2024H 6334H

0 – 0.1 – 0.2 – 0.3

ΔtfbD

Δura3

bat trh5 vacB prp2 0156C 0389C 0591C 0654C 6288C

(i) Survival (N /No)

0798H reductases

TF23

1.0

(ii)

0.1

log10 ratios

6

TFs in group boa4 glcK nusA 1845C tbpA hlx1 0704C phoU sirR imd1 arcR 2126C 2675C 5130H 6126H trh4 0176H 0293H 0147C 0726C 5009H 5050H tfbC hrg 0247C 0511H 1508C 2277H 6287H

8

6239G

2641H AN ND-28

1451C

Gas vesicle kinases

msrA dmsR

TF10 17 Carotenoid biosynthesis

phoU

31

23

71 5 Amino-acid biosynthesis

4

TF24

33 Putative siderophore biosynthesis

TF14

TFGROUP

bc81

bc3

– 0.1

4mM PQ

0.5 0.0

ΔtfbC

Δura3

bc33

0 Δura3

– 0.2 – 0.3

ΔtfbC

6211G 6212G 6213G 6214G 6216G

Figure 4 EGRINOS: environment and gene regulatory influence network for oxidative stress. (A) EGRINOS was constructed by applying the Inferelator algorithm to infer predictive regulatory influences of TFs and EFs (H2O2 and PQ) on biclusters of conditionally co-regulated genes discovered using cMonkey. To simplify visualization of this model, the network was filtered for biclusters with residuals o0.4 and regulatory influence weights 40.1 and restricted to physiological functions relevant to OS. Rectangles with numbers indicate specific biclusters; width of each box is proportional to number of conditions in that bicluster and height is proportional to number of genes. Likewise, width of each arrow is proportional to the strength of the regulatory influence, with red arrows indicating activation and green indicating repression. Circles indicate single TFs or groups of TFs. ‘TF groups’ are averaged expression profiles of their member TFs (see table). Triangles indicate a logical ‘AND’ combination of regulatory influences. A red star indicates that the influence of a TF (red font) within a particular group has been experimentally validated. Corresponding position of the TFgroups in the EGRINos network are indicated with a red border on relevant nodes. (B) Cross-correlation with independently measured protein–DNA interactions by ChIP-chip (Bonneau et al, 2007; Facciotti et al, 2007; Reiss et al, 2008) is consistent with the architecture of the network. TFs in different TFgroups bind promoters of genes within the same bicluster (bc88). Further, these TFs also bind promoters of genes in other biclusters that they are predicted to regulate. The network is drawn as a composite of P–D interactions and statistically inferred influences by Inferelator. Thickness of edges are proportional to significance of promoter binding in ChIP-chip experiments, whereas the activate (red arrows) and repress (green edges) influences are based on Inferelator predictions. (C) TFBd binds the promoters of at least four genes grouped into as many distinct biclusters. Consistent with its predicted influence a knockout in TFBd results in significantly lower transcript levels of these genes at all stages of growth in batch culture. This example validates the EGRINOS predicted regulation by TFBd and illustrates how a TF coordinates transcription of genes involved in different processes during OSR. (D) Finally, the deletion of TFBc, which is predicted to coordinate amino-acid metabolism and iron metabolism genes, results in increased sensitivity to 4 mM PQ [transcription factor-binding sites for the six TFs (indicated with red stars in panel A) are listed in Supplementary Table S9].

(msrA)], cysteine synthase [VNG1301G (cysK)], thioredoxin [VNG1259G (trxB2)], and a ferritin [VNG2443G (dpsA)] (see bc84) suggested a more significant function for this ortholog in providing protection against OS. In agreement with this 8 Molecular Systems Biology 2010

hypothesis, a knockout in sod1, but not VNG1332G (sod2), resulted in hypersensitivity to PQ treatment (Figure 5A and B). However, simultaneous deletions of both orthologs further increased sensitivity to O 2 stress, relative to the Dsod1 mutant & 2010 EMBO and Macmillan Publishers Limited

1 –01

10

10–02 10–03

C

1 10–02

10–06

81

98

ho Δp

ΔV

N

G

G ΔV N

N ΔV

H

H 18

Δu

Δp

07

3 ra

A er

H

2

18

G

d1

00

/s

od

2

1

Δs o

Time (h)

od

50

od

40

Δs

30

a3

20

Δs

10

Δu r

0

10–04

10–08

10–04

0.1

5 mM_H2O2 15 mM_H2O2 25 mM_H2O2

A

OD600

Survival (N i /N o)

Δsod1/sod2 Δsod1 Δsod2 Δura3

er

1

25 mM_H2O2 4 mM PQ

00

B

PQ – 0.125 mM

Δp

A

Survival (N i /N o)

Systems model of oxidative stress response A Kaur et al

A

7

1.6

H2O2(M) per 108 cells

Figure 5 Phenotypic analyses of mutants under PQ and H2O2 stress. (A) Growth analysis of SOD single and double gene knockout mutants in the presence of 0.125 mM PQ. (B) Survival of mutants after a 2 h treatment with either 4 mM PQ or 25 mM H2O2. (C) Survival of mutants after 2 h treatment with 5, 15, or 25 mM H2O2. H. salinarum NRC-1 Dura3 is the parent strain that was used as a control in all experiments.

6

1.4 1.2

5

H2O2(M)_Δura3

0.8 3

H2O2(M)_ΔperA Δura3_OD600 ΔperA_OD600

OD600

1

4

0.6

2

0.4

1

0.2

0 0

20

40

60

80

0 140 160

Time (h) Δura3

ΔperA

Δsod1/sod2

Δura3

C

Integrated absorbance (400–600 nm) (AU)

B

Δura3 Δsod1/sod2 Δsod1 Δsod2

100 10 1 0.1 0

2

4

6

8 10 12 14

Time (d)

Figure 6 Secondary protective mechanisms against OS. (A) Endogenous H2O2 levels during aerobic growth of Dura3 and DperA strains were quantified using the Amplex Red assay. Although H2O2 levels were consistently low throughout growth of the parent strain, as expected the levels were significantly higher in the peroxidase mutant. However, the unexpected growth of this mutant coincided with drop in H2O2 level, suggesting the presence of an alternate mechanism(s) to detoxify H2O2. (B) Defective gas vesicle biogenesis in the peroxidase mutant. To assay for flotation properties, cells were grown to stationary phase at 371C and then left undisturbed (i.e. without shaking) at room temperature on the bench top. Although the parent Dura3 strain floated to the surface, the peroxidase/catalase mutant sank to the bottom of the flask. Microscopy analysis confirmed that gas vesicles (observed as bright refractile bodies in the parent strain) are completely missing in the peroxidase/catalase mutant. (C) The function of carotenoids and other pigments was predicted to be important based on the bleached appearance of stationary phase cultures of the double Dsod1/sod2 mutant. This was confirmed on measuring and calculating a fourfold increase in the rate of degradation of pigments relative to single sod mutants (Dsod1, Dsod2) and the parent Dura3 strain. Pigments were assayed using UV-VIS spectroscopy.

(Figure 5B). Together with the observation that sod2 was upregulated only by treatment with a sub-lethal dose of PQ (Figure 2; Supplementary Figure S3), this suggested that this ortholog provides marginal protection against excessive O 2 stress. However, this does not rule out a more significant & 2010 EMBO and Macmillan Publishers Limited

function for this enzyme under specialized conditions (Valderas and Hart, 2001). Notably, unlike in other organisms (Carlioz and Touati, 1986), neither of the two SODs, individually or together, provided any protection against H2O2 stress (Figure 5B). Molecular Systems Biology 2010 9

Systems model of oxidative stress response A Kaur et al

perA (peroxidase), VNG0798H (a putative ‘dyedecolorizing’ peroxidase), and VNG0018H (a protein of earlier unknown function) offer protection against H2O2 stress and some cross-protection to O 2 Consistent with its anticipated function in scavenging H2O2, deletion of the putative peroxidase/catalase gene perA resulted in poor growth under oxic conditions and hypersensitivity to H2O2 (Figure 5C; Supplementary Figure S6). The onset of growth of the DperA strain after a prolonged lag phase coincided with a corresponding drop in intracellular H2O2 content from B6 to o2 mM (levels observed in typical wildtype cell cultures; all measurements of H2O2 were normalized to 108 cells) (Figure 6A) (Gonzalez-Flecha and Demple, 1997; Seaver and Imlay, 2001). This suggested the presence of additional mechanism(s) to handle peroxide stress. The EGRINOS model suggested two candidate genes: VNG0798H and VNG0018H. VNG0798H, which has a weak borderline match to a protein family signature for dye-decolorizing peroxidases (PF04261, e-value B0.02), was upregulated by both H2O2 and PQ, and co-expressed with carotenoid biosynthesis enzymes, gas vesicle biogenesis proteins, and a putative heat shock protein (bc85) [http://baliga.systemsbiology. net/drupal/content/egrin-oxidative-stress]. Moreover, VNG0798H shares Inferelator-discovered regulatory influences with sod1 and Fe-management proteins [section ‘A unified model for global transcriptional coordination of responses to treatments with H2O2 and PQ’; Figure 4]. Likewise, VNG0018H, whose predicted protein structure (in lieu of absence of a sequencebased match to characterized proteins) matches that of a catalase (PDB: 1cf9A2; Mate et al, 1999) (Bonneau et al, 2004) was upregulated under H2O2 stress. Knockouts in these candidate genes validated their functions in providing partial protection against H2O2 stress (Figure 5C). Although deletion of perA resulted in complete loss of viability at 25 mM H2O2, deletion of VNG0798H and VNG0018H resulted in 99.9 and 62% reduction in survival, respectively (Figure 5C; Supplementary Figure S6). Furthermore, knockouts in perA and VNG0018H also resulted in 54 and 67% loss in survival, respectively, when challenged with 4 mM PQ (Figure 5B; Supplementary Figure S6).

Secondary mechanisms for OS management The co-regulation of genes encoding rhodopsins, carotenoids, and gas vesicles with several known mechanisms to detoxify and manage ROS (e.g. in bc85—http://baliga.systemsbiology. net/drupal/content/egrin-oxidative-stress) suggested important functions for these processes in OS management (Figure 4). Culture characteristics of the perA and SOD mutants (Figure 6B and C) in conjunction with phenotypic analysis and transcriptomics further supported secondary protective functions for these functions and provided clues into their regulatory links to the primary protective mechanisms. (a) Rhodopsins and carotenoids have distinct functions in peroxide and superoxide stress management: Pigments such as carotenoids, which are intermediates in synthesis of the retinal chromophore found in rhodopsins (Kushwaha et al, 1974), are known to act as antioxidants (Di 10 Molecular Systems Biology 2010

Mascio et al, 1991). Consistent with this known function, the transcription of several genes involved in carotenoid biosynthesis [VNG1458G (crtB1), VNG1684G (crtI1), VNG1755G (crtI2), VNG1680G (crtB2), VNG1465G (brp)] and one opsin VNG1467G (bop) were upregulated during and/or after exposure to H2O2 and PQ (Figure 2M and N). This is also consistent with earlier knowledge that transcriptional regulation of these genes is influenced by redox status of the cell (Betlach et al, 1989; Yang and DasSarma, 1990; Shand and Betlach, 1991; Baliga et al, 2001). Furthermore, visual inspection of stationary phase cultures revealed a distinctly bleached appearance of the SOD double mutant—an observation that was confirmed by the fourfold increase in the oxidative degradation rate of pigments during normal growth of this strain relative to the wild-type strain (Figure 6C). The specific mechanism by which carotenoids scavenge ROS is unknown, although it has been speculated that they act through interference with reactions of damaging oxidizing agents (Woodall et al, 1997; Davison et al, 2002). Regardless, an increased rate of pigment bleaching in the SOD double mutant, but not the DperA mutant suggests a specific protective function for carotenoids in defense against O 2 stress. This also partly explains the normal growth characteristics of the single and double SOD mutants under standard laboratory culturing conditions (Supplementary Figure S6). In contrast, depletion of all four rhodopsins [bacteriorhodopsin (bR), halorhodopsin (hR), sensory rhodopsins I and II (sRI and sRII)] as well as two transducers (HtrI and HtrII) in Pho81, a strain constructed earlier for characterizing phototaxis mechanisms (Yao et al, 1994), resulted in increased peroxide toxicity (Figure 5C), but no difference in sensitivity to O 2 (data not shown). In H. salinarum NRC-1, these proteins are involved in light-activated ionpumping (bR and hR) and relocation of cells toward or away from particular wavelengths of radiation by phototaxis [sRI and II, and signal transducers (HtrI and II)]. This is a new function for these integral membrane proteins. The specific mechanisms by which carotenoids and rhodopsins provide protection against H2O2 and PQ stress will require further investigation. (b) Regulation of gas vesicle biogenesis is linked to peroxide stress: H2O2 and PQ treatment induced transcription of genes encoding structural components of gas vesicles. Further, EGRINOS suggested that under OS (especially H2O2 stress; see bc85) at least three of these genes [VNG6236G (gvpG2), VNG6237G (gvpF2), VNG6239G (gvpE2)] are co-induced with the dyp-type peroxidase (VNG0798H), a thioredoxin [VNG5076G (trxA1)], and a putative heat shock protein [VNG1801G (hsp1)]. The induction of purple membrane biogenesis and DMSO fermentation during OS, especially H2O2 treatment (Figure 2) suggests an OS-induced shift in physiology to one that is better suited for low oxygen conditions, so as to minimize the production of additional radicals. Induction of gas vesicle biogenesis under low O2 tension is well known and speculated to be a mechanism used by H. salinarum NRC-1 for scavenging traces of O2 through floatation (Yang and DasSarma, 1990; Schmid et al, 2007). However, in the context of OS, flotation because of vesicle & 2010 EMBO and Macmillan Publishers Limited

Systems model of oxidative stress response A Kaur et al

production would be problematic, as it would increase exposure to O2. Intriguingly, deletion of perA abolished gas vesicle biogenesis resulting in a sinking phenotype (Figure 6B)—a phenomenon also displayed by cyanobacteria that are subjected to OS (Berman-Frank et al, 2004). In H. salinarum NRC-1, this appears to be mediated through post-transcriptional suppression of gas vesicle biogenesis despite increased transcript levels of these genes under conditions of increased H2O2 stress.

Generalized and unique components of OSR It is generally accepted that a significant proportion of cellular responses to diverse types of environmental stress (metals, radiation, starvation, etc.) is shared and constitutes a generalized OS component. We investigated this idea quantitatively by evaluating the degree to which the EGRIN model constructed from responses to perturbations in diverse EFs (six metals, O2, light, UV and g-radiation) was able to predict transcriptional changes induced by H2O2 and PQ treatment. First, using the measured values of just 72 TFs, 9 EFs, and their regulatory influences in the EGRIN model, we predicted transcriptional changes in 80% of all genes in 5000 resampled datasets of 140 experiments each. These datasets were randomly assembled from a pool of 722 new experiments that were not used to construct the model and that did not include any OS experiments (Supplementary Figure S7). This analysis showed the remarkable robustness with which the EGRIN model consistently made accurate global predictions with a mean error (RMSD) of 0.36±0.01. Next, we used the model to make predictions of global gene-expression changes in the 140 experiments (including controls) that specifically investigated transcriptional responses to treatments with H2O2 and PQ. Remarkably, the model predicted global gene-expression changes induced by OS with similar accuracy, suggesting that most regulatory phenomena triggered by OS are also active during responses to other EFs. Historically, such responses that are triggered by diverse environmental stresses are lumped into a loosely defined ‘general stress response.’ Although it might be true that a significant fraction of this general stress response is an indirect consequence of the highly interconnected nature of biological networks, many genes are most certainly triggered by a particular injury or dysfunction. Notwithstanding this observation, a closer investigation revealed OS-specific coordinate regulation of important genes within several regulons that were not observed in responses to any other EFs. Specifically, regulons within 22 of the 67 biclusters within EGRINOS (residuals o0.45) are partially or completely disrupted (R2o0.35) in other environmental stress conditions such as treatments with sub-inhibitory doses of redox-active metals including Cu and Fe (Kaur et al, 2006), a sub-lethal dose of g-radiation (Whitehead et al, 2006), or extreme changes in oxygen tension (Schmid et al, 2007) (Figure 7; Supplementary Table S10). The 288 genes within these 22 biclusters encode functions of detoxification, repair, Fe metabolism, nucleotide metabolism, and oligo/dipeptide ABC transporters. Although this analysis validates the notion that a significant fraction of cellular responses to diverse EFs does indeed constitute a generalized & 2010 EMBO and Macmillan Publishers Limited

component, it also reveals specific regulatory mechanism(s) for the management of excessive OS. The regulon within bc84 is a case in point. Genes within this regulon encode frontline defense mechanisms that are coordinately upregulated by PQ treatment, and to a lesser degree by H2O2 treatment, but not by redox-active metals, g-radiation, or extreme changes in oxygen tension. Co-regulation of Fe-trafficking genes within bc12 is another example of this OS-specific regulatory phenomenon (Figure 7). It will be important to understand natural circumstances that produce such extreme OS, so we can better characterize the functional relevance of such OS-specific regulatory circuits for survival in a hypersaline environment. Especially important will be the dissection of the hierarchy of regulatory networks that coordinate the generalized and specific components of OSR as a function of the degree of OS. The EGRINOS model lends itself to such an analysis by providing conditional interrelationships among TF activities in the form of joint influences on both response components of the OSR.

Summary and conclusion Given the ubiquitous nature of OS, it is not entirely surprising that most organisms have evolved multiple lines of defense— both passive and active. Although many of these mechanisms have been extensively characterized in other organisms, our integrated systems approach has uncovered additional protective mechanisms in H. salinarum (e.g. VNG0798H and VNG0018H). Further, the systems approach has also revealed a structure and hierarchy to the OSR through conditional regulatory associations among various components of the response (Figures 4 and 5). Using this integrated approach, we can begin to synthesize an understanding of the multi-tiered program for management of H2O2 and O 2 stress in H. salinarum NRC-1. Briefly, H2O2 generated by normal metabolism is primarily handled by the peroxidase/catalase PerA; when H2O2 production exceeds detoxification capabilities of PerA, additional protection is provided by VNG0018H, rhodopsins, and VNG0798H, respectively (Figure 5). Uncontrolled accumulation of H2O2 such as on disruption of PerA causes cells to switch to anaerobic physiology, shutdown gas vesicle biogenesis, and sink away from oxygen (Figure 6). Coordinated upregulation of iron trafficking and management systems under severe H2O2 stress (bc12) might also serve to minimize Fenton reaction while keeping with the cellular demand for iron. O 2 stress, on the other hand, is primarily managed by SOD1 with additional marginal protection from SOD2. When challenged with a sub-lethal dose of O 2 , H. salinarum NRC-1 increase the production of carotenoids to take advantage of their secondary function in scavenging ROS. Severe O 2 stress results in the co-regulation of several frontline defense and repair functions including sod1, msrA, and dpsA (bc84) (Figure 3). In addition to such O 2 -specific responses, there was also evidence of crosstalk in responses to H2O2 and O 2 in the coordination of biclusters containing SOD1 (bc84) and VNG0798H (bc6) by the same set of regulatory influences (Figure 4). Likewise, two of the three peroxidases, VNG0018H and perA, also provide significant cross-protection against O 2 Molecular Systems Biology 2010 11

Systems model of oxidative stress response A Kaur et al

OS protection mechanisms. These operational relationships are further extended by the model to other aspects of physiology. For instance, the model reveals that the simultaneous upregulation of genes within bc6 (VNG0798H) and bc84

B

R2

4

H2O2

Paraquat

R =0.778

R =0.648

2

γ-rays

2

2

R =0.32

Oxygen 2

Metals 2

R =0.531 R =0.121

2 0 −2 0

50

100 150 Experimental conditions

200

Normalized log10 ratio

Bicluster 37 (21 genes): ribosomal proteins, RNAP

4

H2O2

Paraquat

R =0.83

R =0.89

2

γ-rays

2

2

R =0.402

Oxygen 2

Metals 2

R =0.773 R =0.606

2 0 −2 0

50

100 150 Experimental conditions

200

Bicluster 12 (21 genes): Fe transport, siderophore biosynthesis, Trx reductase

Normalized log10 ratio

65 56 28 91 31 85 81 50 69 58 70 3 86 24 100 73 93 30 45 54 72 98 77 5 48 53 33 18 19 37 55 97 80 15 29 43 39 44 12 71 88 17 79 52 23 38 40 20 94 78 26 21 59 46 6 57 61 22 51 13 27 87 84 34 74 76 9

Normalized log10 ratio

Bicluster 31 (25 genes): gas vesicle biogenesis, K uptake,

4

H2O2

Paraquat

R =0.586

R =0.404

2

γ-rays

2

2

R =0.312

Oxygen 2

Metals 2

R =0.318 R =0.112

2 0 −2 0

50

100 150 Experimental conditions

200

Bicluster 84 (23 genes): SOD (sod1), msrA, Cys synthase, protease, chaperone, Trx, Trx reductase, ferritin

Normalized log10 ratio

A

H2O2 PQ O2 γ rays Metals

stress (Figure 5). Although this is to be expected as ROS are readily inter-converted such as by dismutation of O 2 to H2O2, these network connections have also provided insights into the operational relationships in the crosstalk among two distinct

4

H2O2

Paraquat

R =0.33

R =0.748

2

γ-rays

2

2

R =0.035

Oxygen 2

Metals 2

R =0.025 R =0.062

2 0 −2

100 150 200 Experimental conditions : OS-specific biclusters R 2 < 0.35 in all environmental conditions other than H2O2 and PQ treatments

0.0

log10 ratio

C

0

50

0.6

0.6

Constant stress Recovery from stress

sod1 msrA

Co/Cu/Fe γ-rays Mn

O2

0.4 0.2 0 – 0.2 – 0.4 H2O2

Ni

PQ

Zn

Experimental conditions

12 Molecular Systems Biology 2010

& 2010 EMBO and Macmillan Publishers Limited

Systems model of oxidative stress response A Kaur et al

(sod1, msrA, dpsA) under certain OS conditions is coordinated with downregulation of cobalamin synthesis and oxidative phosphorylation (bc91) in a biologically meaningful manner (Figure 4). We have validated some aspects of the architecture of this OSR network by confirming physical protein–DNA interactions of six TFs with promoters of genes they were predicted to influence by EGRINOS (Figure 4B). Furthermore, we have also shown that deleting the TFs results in decreased transcript levels of predicted target genes and that disruption of this control can lead to decreased survival rate under OS (Figure 4C and D). It is notable that the five GTFs whose functions we have validated cannot directly sense ROS. This illustrates the importance of the EGRINOS architecture and the functions GTFs have therein to globally coordinate various processes during OSR. This systems level insight would not be possible without integration of diverse global datasets. Finally, by comparing across active regulatory programs under diverse environmental stresses, we have quantitatively determined processes that are specifically triggered by extreme OS (Figure 7).

Materials and methods Organism and growth conditions H. salinarum NRC-1 (wild type), H. salinarum NRC-1 Dura3 (parent for knockout strains), and all gene knockout strains were grown in complex medium (CM: 250 g/l NaCl, 20 g/l MgSO4, 2 g/l KCl, 3 g/l sodium citrate, 10 g/l Oxoid brand bacteriological peptone) at 371C and 220 r.p.m. shaking in Innova9400 incubator (New Brunswick). CM was supplemented with 50 mg/l uracil for strains constructed from a Dura3 background. Gene deletion mutants were constructed using a two-step in-frame gene replacement strategy as described earlier (Kaur et al, 2006) and rhodopsin-deficient strain, Pho81, has been characterized earlier (Yao et al, 1994).

Growth and survival assay All strains were grown in CM with continuous shaking (200 r.p.m.) at 371C until late log phase (ODB0.8) and then further diluted into fresh medium to an optical density of 0.05. These cultures were then transferred to Honeycomb plate wells of Bioscreen C (Growth Curves USA, Piscataway, NJ). Different amounts of PQ (0, 0.125, 0.25, and 0.5 mM—final concentrations) or H2O2 (0, 4, 5, 6, and 7 mM—final concentrations) were added into individual wells. Samples were incubated at 371C with continuous shaking and growth was monitored for 48 h in Bioscreen. For survival assay, H. salinarum NRC-1 was exposed to either 0–100 mM H2O2 or 0–8 mM PQ, whereas knockout strains were exposed to only 25 mM H2O2 or 4 mM PQ for 2-h treatment. Cell viability was determined by counting colonies on agar plates. Sensitivity of Dura3, DperA, DVNG0798H, and Pho81 were also tested with different concentrations of H2O2 (5, 15, 25 mM). Three independent measurements were made for each mutant.

Exposure to H2O2 and PQ for RNA preparation and microarray analysis Two time courses were run to determine the transcriptional responses to (1) constant stress and (2) recovery of H. salinarum NRC-1 to H2O2 and PQ. Mid-log phase cultures grown in flasks were treated with sublethal concentrations of 25 mM H2O2 or 4 mM PQ and incubated at 371C with shaking for up to 240 min. During constant stress, culture aliquots (B4 ml) were collected over a time course (1, 5, 10, 20, 40, 80, and 160 min), centrifuged (16 000 g, 90 s), and flash frozen. For recovery, cells were first treated for 2 h with either 25 mM H2O2 or 4 mM PQ, recovered by centrifugation, washed, and re-suspended in CM. Cultures were returned to the incubator with shaking and samples were taken at 0, 10, 20, 30, 40, 50, 60, 120, and 240 min and processed as described earlier. Analysis of temporal transcriptional changes (1, 0, 5, 10, 20, 40, 80, 160, and 320 m) to sub-inhibitory concentrations of PQ (0.25 mM) was performed using the BioFlo110 modular bench-top fermentor (New Brunswick Scientific). Total RNA was prepared using the Absolutely RNA kit (Stratagene) according to manufacturer’s instructions. Microarray slide fabrication, labeling with Alexa547 and Alexa647 dyes (Molecular Probes and Kreatech BV), hybridization, and washing were performed as described earlier (Baliga et al, 2004). Raw intensity signals were processed and resultant data were median normalized and evaluated for statistical significance of differential expression with significance of microarray (SAM) and variability and error estimates (VERA) algorithms (Ideker et al, 2000). In constant datasets, mRNA log10 ratios were normalized to first condition (control) and data was filtered based on fold change 40.05, l¼15 in at least one condition. In recovery sets, data was filtered (fold change 40.05, l¼15) and compared with temporal changes in control experiments. A total of B1400 genes met these criteria. Genes were further classified into different functional groups according to Kyoto Encyclopedia of Genes and Genomes, PFAM (http://pfam.sanger. ac.uk/), and clusters of orthologous groups (http://www.ncbi.nih. gov/COG). Analysis of microarray results was performed using hierarchical clustering and Dynamic Regulatory Events Miner (Ernst et al, 2007), which uses a hidden Markov model to construct dynamic models for condition-specific gene regulation. The complete microarray dataset is available at http://www.ncbi.nlm.nih.gov/geo/query/ acc.cgi?acc¼GSE17515.

H2O2 production during aerobic growth H2O2 production was measured with Amplex Red H2O2 assay (Molecular Probes—A22188). Late log phase cultures of Dura3, DperA strains were diluted 1:100 in fresh CM and grown for six days. H2O2 levels were measured along the growth curve at different time points (21, 29, 36, 48, 72, and 144 h). The rate of H2O2 was normalized to the number of cells. The relation between OD600 and cell number was calculated by counting cells in a flow cytometer.

Pigment analysis Strains (Dura3, Dsod1, Dsod2, Dsod1/sod2) were grown in CM media at 371C until late stationary phase. Aliquots (1 ml) were withdrawn from cultures at different time points (B24 h intervals), centrifuged at 16 000 g for 2 min. Pigments were extracted from cell pellets with 90% cold acetone and analyzed with UV-VIS spectrophotometer (Beckman

Figure 7 Coordination of frontline defense mechanisms under extreme OS. (A) Degree of co-regulation of genes within 67 biclusters (residuals o0.45) in diverse environmental stress conditions was evaluated by calculating correlations (R2) among their transcript level changes in those experiments. The R2-values were hierarchically clustered to identify the generalized and specific components of OSR. (B) Pearson’s correlation (R) among mean and variance-normalized transcriptional changes for four representative biclusters are shown. Biclusters 31 and 37 are examples of the generalized stress response component of OSR with co-regulation across most environmental conditions. In contrast, co-regulation of genes within biclusters 12 and 84 is much more significant under severe OS conditions. For instance, genes of bc84 are co-regulated under severe OS resulting from treatment with sub-lethal dose PQ and to a lesser degree H2O2, but not on irradiation with a sub-lethal dose of g-irradiation (Whitehead et al, 2006), subjecting cells to sudden and extreme changes in O2 tension (Schmid et al, 2007) or treatment with sub-inhibitory dose of transition metals (Kaur et al, 2006). Genes within bc12, on the other hand, were better co-regulated on treatment with H2O2. Genes within both biclusters (especially bc84) encode important functions associated with frontline defense mechanisms that provide protection against OS (see Figure 3 and http://baliga.systemsbiology.net/drupal/content/ egrin-oxidative-stress for characteristics of biclusters). (C) A specific example of OS-specific co-regulation: two genes of bc84, sod1 and msrA, are co-induced under severe OS—especially with PQ treatment—but not coordinately controlled by any other stressful environmental conditions.

& 2010 EMBO and Macmillan Publishers Limited

Molecular Systems Biology 2010 13

Systems model of oxidative stress response A Kaur et al

Coulter—DU800). Absorption spectra were recorded and analyzed between 400 and 600 nm.

Gas vesicles visualization Aliquots were harvested at various time points in Dura3 and DperA strains and further diluted to an OD600¼0.1–0.2. Cells were then fixed by the addition of 0.25% formaldehyde (final concentration) and imaged with phase contrast microscopy.

Calculation of significance of enrichment List of oxygen responsive genes was generated with principal component analysis of oxygen microarray experiment from earlier study (Schmid et al, 2007), which resulted in two categories of genes; one correlated (oxic genes—105) and other anti-correlated (anoxic genes—110) with oxygen. P-values were computed for enrichment of both oxic and anoxic genes among all OS responsive genes based on hyper-geometric distribution.

Motif search Motifs conserved for various functions such as redox (CX2S), Fe–S (CX2CX2CX3C and CX3CX2C), conserved cysteines residues of oxyR protein (CX8C), and metal containing cysteine residues (CX2C) along with PAS (PF00989, PF00785), GAF (PF01590), and MCP(PF00015)signaling domain were identified from literature (Zheng and Storz, 2000; Fomenko and Gladyshev, 2002; Paget and Buttner, 2003; Fontecave, 2006). MatLab scripts were written to search the H. salinarum NRC-1 proteome for matches to one or more of these conserved motifs (Supplementary Figure S4; Supplementary Table S5).

Accession sites Access to modeling results online @: http://baliga.systems biology.net/drupal/content/egrin-oxidative-stress. GEO accession number for microarray data: GSE17515 (http:// www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc¼GSE17515).

Supplementary information Supplementary information is available at the Molecular Systems Biology website (http://www.nature.com/msb).

Acknowledgements This work was supported by grants from NIH (P50GM076547 and 1R01GM077398-01A2), DoE (DOE SFA: ENIGMA: DE-FG02-07ER64327 and DE-FG02-07ER64327), and NSF (EF-0313754, EIA-0220153, MCB0425825, DBI-0640950) to NSB; AFOSR (FA95500710158) to JDR and NASA (NNG05GN58G) to NSB and JDR. We thank Dan Tenenbaum for help in construction of the webpage. We also thank John Spudich for providing us with the rhodopsin-deficient strain Pho81.

Conflict of interest The authors declare that they have no conflict of interest.

References Andrews SC, Robinson AK, Rodriguez-Quinones F (2003) Bacterial iron homeostasis. FEMS Microbiol Rev 27: 215–237 Apel K, Hirt H (2004) Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol 55: 373–399 14 Molecular Systems Biology 2010

Baliga NS, Bjork SJ, Bonneau R, Pan M, Iloanusi C, Kottemann MC, Hood L, DiRuggiero J (2004) Systems level insights into the stress response to UV radiation in the halophilic archaeon Halobacterium NRC-1. Genome Res 14: 1025–1035 Baliga NS, Kennedy SP, Ng WV, Hood L, DasSarma S (2001) Genomic and genetic dissection of an archaeal regulon. Proc Natl Acad Sci USA 98: 2521–2525 Berman-Frank I, Bidle KD, Haramaty L, Falkowski PG (2004) The demise of the marine cyanobacterium, Trichodesmium spp., via an autocatalyzed cell death pathway. Limnol Oceanogr 49: 997–1005 Betlach MC, Shand RF, Leong DM (1989) Regulation of the bacterioopsin gene of a halophilic archaebacterium. Can J Microbiol 35: 134–140 Bonneau R, Baliga NS, Deutsch EW, Shannon P, Hood L (2004) Comprehensive de novo structure prediction in a systems-biology context for the archaea Halobacterium sp. NRC-1. Genome Biol 5: R52 Bonneau R, Facciotti MT, Reiss DJ, Schmid AK, Pan M, Kaur A, Thorsson V, Shannon P, Johnson MH, Bare JC, Longabaugh W, Vuthoori M, Whitehead K, Madar A, Suzuki L, Mori T, Chang DE, Diruggiero J, Johnson CH, Hood L et al (2007) A predictive model for transcriptional control of physiology in a free living cell. Cell 131: 1354–1365 Bonneau R, Reiss DJ, Shannon P, Facciotti M, Hood L, Baliga NS, Thorsson V (2006) The Inferelator: an algorithm for learning parsimonious regulatory networks from systems-biology data sets de novo. Genome Biol 7: R36 Boyd JM, Drevland RM, Downs DM, Graham DE (2009) Archaeal ApbC/Nbp35 homologs function as iron-sulfur cluster carrier proteins. J Bacteriol 191: 1490–1497 Brioukhanov AL, Netrusov AI, Eggen RI (2006) The catalase and superoxide dismutase genes are transcriptionally up-regulated upon oxidative stress in the strictly anaerobic archaeon Methanosarcina barkeri. Microbiology 152(Part 6): 1671–1677 Cabiscol E, Tamarit J, Ros J (2000) Oxidative stress in bacteria and protein damage by reactive oxygen species. Int Microbiol 3: 3–8 Carlioz A, Touati D (1986) Isolation of superoxide dismutase mutants in Escherichia coli: is superoxide dismutase necessary for aerobic life? EMBO J 5: 623–630 Causton HC, Ren B, Koh SS, Harbison CT, Kanin E, Jennings EG, Lee TI, True HL, Lander ES, Young RA (2001) Remodeling of yeast genome expression in response to environmental changes. Mol Biol Cell 12: 323–337 Coulson RM, Touboul N, Ouzounis CA (2007) Lineage-specific partitions in archaeal transcription. Archaea 2: 117–125 Davison PA, Hunter CN, Horton P (2002) Overexpression of betacarotene hydroxylase enhances stress tolerance in Arabidopsis. Nature 418: 203–206 Demple B (1997) Study of redox-regulated transcription factors in prokaryotes. Methods 11: 267–278 Di Mascio P, Murphy ME, Sies H (1991) Antioxidant defense systems: the role of carotenoids, tocopherols, and thiols. Am J Clin Nutr 53 (1 Suppl): 194S–200S Ernst J, Vainas O, Harbison CT, Simon I, Bar-Joseph Z (2007) Reconstructing dynamic regulatory maps. Mol Syst Biol 3: 74 Facciotti MT, Reiss DJ, Pan M, Kaur A, Vuthoori M, Bonneau R, Shannon P, Srivastava A, Donohoe SM, Hood LE, Baliga NS (2007) General transcription factor specified global gene regulation in archaea. Proc Natl Acad Sci USA 104: 4630–4635 Farr SB, Kogoma T (1991) Oxidative stress responses in Escherichia coli and Salmonella typhimurium. Microbiol Rev 55: 561–585 Finkel T (2005) Radical medicine: treating ageing to cure disease. Nat Rev Mol Cell Biol 6: 971–976 Fomenko DE, Gladyshev VN (2002) CxxS: fold-independent redox motif revealed by genome-wide searches for thiol/disulfide oxidoreductase function. Protein Sci 11: 2285–2296 Fontecave M (2006) Iron-sulfur clusters: ever-expanding roles. Nat Chem Biol 2: 171–174

& 2010 EMBO and Macmillan Publishers Limited

Systems model of oxidative stress response A Kaur et al

Gonzalez-Flecha B, Demple B (1997) Homeostatic regulation of intracellular hydrogen peroxide concentration in aerobically growing Escherichia coli. J Bacteriol 179: 382–388 Grant CM (2001) Role of the glutathione/glutaredoxin and thioredoxin systems in yeast growth and response to stress conditions. Mol Microbiol 39: 533–541 Green J, Paget MS (2004) Bacterial redox sensors. Nat Rev Microbiol 2: 954–966 Hassan HM, Fridovich I (1979) Paraquat and Escherichia coli. Mechanism of production of extracellular superoxide radical. J Biol Chem 254: 10846–10852 Ideker T, Thorsson V, Siegel AF, Hood LE (2000) Testing for differentially-expressed genes by maximum-likelihood analysis of microarray data. J Comput Biol 7: 805–817 Imlay JA (2003) Pathways of oxidative damage. Annu Rev Microbiol 57: 395–418 Imlay JA (2006) Iron-sulphur clusters and the problem with oxygen. Mol Microbiol 59: 1073–1082 Imlay JA (2008) Cellular defenses against superoxide and hydrogen peroxide. Annu Rev Biochem 77: 755–776 Joshi P, Dennis PP (1993) Characterization of paralogous and orthologous members of the superoxide dismutase gene family from genera of the halophilic archaebacteria. J Bacteriol 175: 1561–1571 Kaur A, Pan M, Meislin M, Facciotti MT, El-Geweley R, Baliga NS (2006) A systems view of haloarchaeal strategies to withstand stress from transition metals. Genome Res 16: 841–854 Kawakami R, Sakuraba H, Kamohara S, Goda S, Kawarabayasi Y, Ohshima T (2004) Oxidative stress response in an anaerobic hyperthermophilic archaeon: presence of a functional peroxiredoxin in Pyrococcus horikoshii. J Biochem 136: 541–547 Kushwaha SC, Gochnauer MB, Kushner DJ, Kates M (1974) Pigments and isoprenoid compounds in extremely and moderately halophilic bacteria. Can J Microbiol 20: 241–245 Lee JW, Helmann JD (2006) The PerR transcription factor senses H2O2 by metal-catalysed histidine oxidation. Nature 440: 363–367 Limauro D, Pedone E, Galdi I, Bartolucci S (2008) Peroxiredoxins as cellular guardians in Sulfolobus solfataricus: characterization of Bcp1, Bcp3 and Bcp4. FEBS J 275: 2067–2077 Limauro D, Pedone E, Pirone L, Bartolucci S (2006) Identification and characterization of 1-Cys peroxiredoxin from Sulfolobus solfataricus and its involvement in the response to oxidative stress. FEBS J 273: 721–731 Liochev SI, Fridovich I (2007) The effects of superoxide dismutase on H2O2 formation. Free Radic Biol Med 42: 1465–1469 Loiseau L, Ollagnier-de-Choudens S, Nachin L, Fontecave M, Barras F (2003) Biogenesis of Fe-S cluster by the bacterial Suf system: SufS and SufE form a new type of cysteine desulfurase. J Biol Chem 278: 38352–38359 Mate MJ, Sevinc MS, Hu B, Bujons J, Bravo J, Switala J, Ens W, Loewen PC, Fita I (1999) Mutants that alter the covalent structure of catalase hydroperoxidase II from Escherichia coli. J Biol Chem 274: 27717–27725 Metayer S, Seiliez I, Collin A, Duchene S, Mercier Y, Geraert PA, Tesseraud S (2008) Mechanisms through which sulfur amino acids control protein metabolism and oxidative status. J Nutr Biochem 19: 207–215 Moore CM, Helmann JD (2005) Metal ion homeostasis in Bacillus subtilis. Curr Opin Microbiol 8: 188–195 Oliver AE, Leprince O, Wolkers WF, Hincha DK, Heyer AG, Crowe JH (2001) Non-disaccharide-based mechanisms of protection during drying. Cryobiology 43: 151–167 Outten FW (2007) Iron-sulfur clusters as oxygen-responsive molecular switches. Nat Chem Biol 3: 206–207 Paget MS, Buttner MJ (2003) Thiol-based regulatory switches. Annu Rev Genet 37: 91–121 Pahlman AK, Granath K, Ansell R, Hohmann S, Adler L (2001) The yeast glycerol 3-phosphatases Gpp1p and Gpp2p are required for glycerol biosynthesis and differentially involved in the cellular

& 2010 EMBO and Macmillan Publishers Limited

responses to osmotic, anaerobic, and oxidative stress. J Biol Chem 276: 3555–3563 Perrone GG, Tan SX, Dawes IW (2008) Reactive oxygen species and yeast apoptosis. Biochim Biophys Acta 1783: 1354–1368 Reindel S, Schmidt CL, Anemuller S, Matzanke BF (2006) Expression and regulation pattern of ferritin-like DpsA in the archaeon Halobacterium salinarum. Biometals 19: 19–29 Reiss DJ, Baliga NS, Bonneau R (2006) Integrated biclustering of heterogeneous genome-wide datasets for the inference of global regulatory networks. BMC Bioinformatics 7: 280 Reiss DJ, Facciotti MT, Baliga NS (2008) Model-based deconvolution of genome-wide DNA binding. Bioinformatics 24: 396–403 Scheibner M, Hulsdau B, Zelena K, Nimtz M, de Boer L, Berger RG, Zorn H (2008) Novel peroxidases of Marasmius scorodonius degrade beta-carotene. Appl Microbiol Biotechnol 77: 1241–1250 Schmid AK, Reiss DJ, Kaur A, Pan M, King N, Van PT, Hohmann L, Martin DB, Baliga NS (2007) The anatomy of microbial cell state transitions in response to oxygen. Genome Res 17: 1399–1413 Seaver LC, Imlay JA (2001) Alkyl hydroperoxide reductase is the primary scavenger of endogenous hydrogen peroxide in Escherichia coli. J Bacteriol 183: 7173–7181 Shand RF, Betlach MC (1991) Expression of the bop gene cluster of Halobacterium halobium is induced by low oxygen tension and by light. J Bacteriol 173: 4692–4699 Shukla HD (2006) Proteomic analysis of acidic chaperones, and stress proteins in extreme halophile Halobacterium NRC-1: a comparative proteomic approach to study heat shock response. Proteome Sci 4: 6 Stadtman ER (2006) Protein oxidation and aging. Free Radic Res 40: 1250–1258 Stadtman ER, Moskovitz J, Levine RL (2003) Oxidation of methionine residues of proteins: biological consequences. Antioxid Redox Signal 5: 577–582 Storz G, Imlay JA (1999) Oxidative stress. Curr Opin Microbiol 2: 188–194 Sugano Y, Muramatsu R, Ichiyanagi A, Sato T, Shoda M (2007) DyP, a unique dye-decolorizing peroxidase, represents a novel heme peroxidase family: ASP171 replaces the distal histidine of classical peroxidases. J Biol Chem 282: 36652–36658 Takahashi Y, Tokumoto U (2002) A third bacterial system for the assembly of iron-sulfur clusters with homologs in archaea and plastids. J Biol Chem 277: 28380–28383 Temple MD, Perrone GG, Dawes IW (2005) Complex cellular responses to reactive oxygen species. Trends Cell Biol 15: 319–326 Thorpe GW, Fong CS, Alic N, Higgins VJ, Dawes IW (2004) Cells have distinct mechanisms to maintain protection against different reactive oxygen species: oxidative-stress-response genes. Proc Natl Acad Sci USA 101: 6564–6569 Toledano MB, Kumar C, Le Moan N, Spector D, Tacnet F (2007) The system biology of thiol redox system in Escherichia coli and yeast: differential functions in oxidative stress, iron metabolism and DNA synthesis. FEBS Lett 581: 3598–3607 Touati D (2000) Iron and oxidative stress in bacteria. Arch Biochem Biophys 373: 1–6 Tusher VG, Tibshirani R, Chu G (2001) Significance analysis of microarrays applied to the ionizing radiation response. Proc Natl Acad Sci USA 98: 5116–5121 Valderas MW, Hart ME (2001) Identification and characterization of a second superoxide dismutase gene (sodM) from Staphylococcus aureus. J Bacteriol 183: 3399–3407 Varghese S, Wu A, Park S, Imlay KR, Imlay JA (2007) Submicromolar hydrogen peroxide disrupts the ability of Fur protein to control free-iron levels in Escherichia coli. Mol Microbiol 64: 822–830 Whitehead K, Kish A, Pan M, Kaur A, Reiss DJ, King N, Hohmann L, DiRuggiero J, Baliga NS (2006) An integrated systems approach for understanding cellular responses to gamma radiation. Mol Syst Biol 2: 47 Woodall AA, Lee SW, Weesie RJ, Jackson MJ, Britton G (1997) Oxidation of carotenoids by free radicals: relationship Molecular Systems Biology 2010 15

Systems model of oxidative stress response A Kaur et al

between structure and reactivity. Biochim Biophys Acta 1336: 33–42 Yang CF, DasSarma S (1990) Transcriptional induction of purple membrane and gas vesicle synthesis in the archaebacterium Halobacterium halobium is blocked by a DNA gyrase inhibitor. J Bacteriol 172: 4118–4121 Yao VJ, Spudich EN, Spudich JL (1994) Identification of distinct domains for signaling and receptor interaction of the sensory rhodopsin I transducer, HtrI. J Bacteriol 176: 6931–6935 Zheng M, Storz G (2000) Redox sensing by prokaryotic transcription factors. Biochem Pharmacol 59: 1–6

16 Molecular Systems Biology 2010

Zheng M, Wang X, Templeton LJ, Smulski DR, LaRossa RA, Storz G (2001) DNA microarray-mediated transcriptional profiling of the Escherichia coli response to hydrogen peroxide. J Bacteriol 183: 4562–4570

Molecular Systems Biology is an open-access journal published by European Molecular Biology Organization and Nature Publishing Group. This work is licensed under a Creative Commons Attribution-Noncommercial-Share Alike 3.0 Unported License.

& 2010 EMBO and Macmillan Publishers Limited